• Aucun résultat trouvé

Perturbations of quadratic hamiltonian systems with symmetry

N/A
N/A
Protected

Academic year: 2022

Partager "Perturbations of quadratic hamiltonian systems with symmetry"

Copied!
41
0
0

Texte intégral

(1)

A NNALES DE L ’I. H. P., SECTION C

E MIL I VANOV H OROZOV

I LIYA D IMOV I LIEV

Perturbations of quadratic hamiltonian systems with symmetry

Annales de l’I. H. P., section C, tome 13, n

o

1 (1996), p. 17-56

<http://www.numdam.org/item?id=AIHPC_1996__13_1_17_0>

© Gauthier-Villars, 1996, tous droits réservés.

L’accès aux archives de la revue « Annales de l’I. H. P., section C » (http://www.elsevier.com/locate/anihpc) implique l’accord avec les condi- tions générales d’utilisation (http://www.numdam.org/conditions). Toute uti- lisation commerciale ou impression systématique est constitutive d’une infraction pénale. Toute copie ou impression de ce fichier doit conte- nir la présente mention de copyright.

Article numérisé dans le cadre du programme Numérisation de documents anciens mathématiques

http://www.numdam.org/

(2)

Perturbations of quadratic

Hamiltonian systems with symmetry 1

Emil Ivanov HOROZOV

Iliya

Dimov ILIEV

University of Sofia, Faculty of Mathematics and Informatics, 5, J. Bourchier blvd., 1126 Sofia, Bulgaria.

Institute of Mathematics, Bulgarian Academy of Sciences, P.O. Box 373, 1090 Sofia, Bulgaria.

Vol. 13, 1, 1996, p. 17-56 Analyse non linéaire

ABSTRACT. - It is

proved

that

arbitrary quadratic perturbations

of

quadratic

Hamiltonian

systems

in the

plane possessing

central

symmetry

can

produce

at most two limit

cycles.

Key words: Bifurcation, limit cycle, quadratic vector field, Abelian integrals.

RESUME. - Nous prouvons que les

perturbations quadratiques

arbitraires des

systemes

hamiltoniens

quadratiques

et

symetrie

sur la

plane peut produire

au

plus

deux

cycles

limites.

0. INTRODUCTION

In this paper we prove the

following

result:

1 Research partially supported by Grants MM - 402/94 and MM - 410/94 of the NSF of

Bulgaria.

Annales de l’lnstitut Henri Poincaré - Analyse non linéaire - 0294-1449 Vol. 13/96/01/$ 4.00/@ Gauthier-Villars

(3)

THEOREM 1. - Let H be a

generic

cubic Hamiltonian with a central symmetry :

H(-x, -~) _ -H(x, y).

Then any

quadratic perturbation of

the

corresponding

Hamiltonian system

has at most two limit

cycles for E

small

enough.

Our interest in

systems (0.1)

was motivated

by

the

project

to obtain an

explicit

and

sharp

bound

c(2)

for the number of limit

cycles

in

quadratic

vector fields which are close to Hamiltonian ones or at least to obtain an

explicit sharp

bound for the so called Hilbert-Arnold

problem [2]

for n = 2

(see [9], [12], [15], [16], [27]

for results in this

direction).

The Hilbert- Arnold

problem

is formulated as follows: find an upper bound

Z(n)

for

the number of isolated zeros of the

integral

where

b ( h)

is the oval

(compact

connected

component)

of the

algebraic

curve

H(x, y)

= hand

deg H

= n +

1, deg f ,

g = n

(or

more

generally,

find an upper bound

Z(n, m) provided max (deg f , deg g)

=

m).

Integrals

of this

type

in studies of limit

cycles

were introduced

by Pontrjagin [23].

The first nontrivial case was considered

by Bogdanov [6].

The existence of an upper bound

Z(n)

was established

by

A. Varchenko

[25]

and

Khovanskii [20].

But

making

this bound

explicit

seems to be a difficult

problem. Quite recently,

Yu.

Il’ yashenko

and S. Yakovenko

[19] proved

a

double

exponential

estimate for

Z(n, m) provided H

is a fixed Hamiltonian from some dense set of

generic

Hamiltonians:

Z(n, m) 22~~m~ . Although

this bound is far from the

expected (polynomial)

one, up to now this is the best known result.

More

generally,

one can consider

perturbations

of

integrable systems

instead of Hamiltonian ones. In the

quadratic

case, an essential

part

of the

problem

also consists

(see [28])

in

counting

the isolated zeros of certain

(not necessarily Abelian) integrals. Zol~dek

has

proved [28]

that in

quadratic perturbations

of Lotka-Volterra

systems

with a center the

corresponding integrals

will have at most two zeros.

Among

the

remaining

three classes of

quadratic systems

with a center, most of results have been obtained for the Hamiltonian case

(see

e.g.

[17]).

Almost

nothing

is known about the other two

classes,

where at least three limit

cycles

can appear.

Annales de l’Institut Henri Poincaré - Analyse non linéaire

(4)

In a recent paper

[ 16]

we found that for

generic

cubic Hamiltonians with three saddles and one centre the exact value for both

c(2)

and

Z(2)

is two.

For the two

remaining

basic classes of cubic Hamiltonians

(namely

those

with one saddle and one centre and

respectively

with both two saddles and

centres)

the

problem

is still open. Our

conjecture

is that the exact bound for all

generic

cases is two

(the

same bound was

conjectured

in

[28]).

The

symmetric

Hamiltonians form a co-dimension one subset within the class of Hamiltonians

having

two saddles and two centres. For many reasons it seems to us that the

symmetric

case is the one

allowing

to be treated

comparatively

easier at least as far as the

computations

are concerned.

In the statement of Theorem 1 as well as in the above comments we use the notion of

"generic"

Hamiltonian. There are several definitions of this notion even for

quadratic

Hamiltonian vector fields.

Throughout

this

paper we say the cubic H is

generic

if the

corresponding

Hamiltonian

vector field dH = 0 has a centre and does not

belong simultaneously

to any of the other

integrable

classes of

quadratic

vector fields

(as

listed

in

[28]).

The latter condition has a

simple geometrical interpretation:

non-

generic

are

exactly

those Hamiltonians for which the level curves in suitable coordinates have an axis of

symmetry.

On its hand

evidently

the central

symmetry

of H

geometrically

causes dH = 0 to be

centrosymmetric

with

respect

to a suitable

point (xo,

which we without loss of

generality

have chosen to coincide with the

origin.

The

general

case

H(xo - x, y) = -.H(~ -

is dealt with a mere translation.

One can

easily

check that after eventual linear

change

of the variables

each cubic Hamiltonian with a central

symmetry

and

having

a centre can

be written in the form

Then

"generic"

in this case

means ~

0. Selected level curves of H for

/1 =

0, _~c

E

(0,1)

and /1 = 1 are drawn in

Fig.

1.

System (0.1 )

has attracted the interest of several other authors

[4], [9], [21].

In

particular

Bamon

[4]

found a value of /1

(~c

=

(4 3 -1) / (4 3 + 1)

=

0,227....) and

a suitable

perturbation

for which around one of the foci a

saddle

loop

and at least one limit

cycle

can coexist. Drachman et al.

[9]

proved

that

for

close to 1 and for a

specific perturbation ( f , g)

the

system (0.1)

has two limit

cycles

around one of the foci and has no limit

cycle

around the other one.

(In [4], [9]

a different coordinate

system

was

used).

Our paper covers all

generic symmetric

Hamiltonians and all small

quadratic perturbations showing

that the

sharp

estimate of the total number of limit

cycles

in

(0.1)

is two. Moreover from our

analysis

it follows that

Vol.

(5)

all numbers 2 and

positions

of the limit

cycles permitted by

the

general theory

of

quadratic systems [7]

can be realized. It is also easy to describe all bifurcations when the

parameters

of

( f , g)

vary. Some of these results

(for special perturbations)

were announced in

[21] ] but apparently

without

proofs.

In this paper we do not consider the

degenerate

case = 0 which

requires

an

analysis

up to a fourth order in c. This case as well as the other

(nonsymmetric) degenerate

cases are an

object

of another research

(in progress).

Perhaps

it is worth

noticing

that

system (0.1)

falls into class

IIIa~o

of the

Chinese classification

[26].

More

precisely

this class consists of the

systems

When

b = l = m = b = 0, n ~

0 the

system (0.4)

is Hamiltonian and

centrosymmetric

with

respect

to the

point ( - 2a , 2n ) .

Then a

simple

corollary

from Theorem 1 is the

following

one:

COROLLARY 1. -

In I ~

0 and small

b, l,

m;

b,

system

(0.4)

has no more than two limit

cycles.

The

techniques

we use to

get

the results of the

present

paper is a combination of the

techniques

from

[16]

with some ideas from

[11].

In

particular

we

extensively

use the notion and the

properties

of the centroid

curve.

(A

centroid curve is formed

by

the mass centres of a continuous

family

of ovals within the level curves of

H.)

There are two centroid

curves in the considered case,

corresponding

to two

periodic

annuli. In our

study

the

elementary

fact

applies

that each intersection

point

of the zero

divergence

line in

(0.1)

with a centroid curve

corresponds

to a zero of

I ( h)

and hence to a limit

cycle. Using

the mutual

position

of the two centroid

curves, Theorem 1 is

easily

seen to be a consequence of the fact that in the

generic

case each of the centroid curves has a non-zero curvature.

A

convexity argument

is

quite

natural in situations where the

expected

number of

cycles

is two.

Incidentally,

a

convexity

idea has been used

by Il’ yashenko [18]

in his paper about Abelian

integrals arising

in

symmetric perturbations

of Hamiltonian

systems

with

symmetry

of order two.

A more detailed sketch of our construction which we believe will

help

. the reader to overcome easier the rather technical

proof

is

given

in the

next section.

Annales de l’lnstitut Henri Poincaré - Analyse non linéaire

(6)

1. PRELIMINARIES AND OUTLINE

OF

THE PROOF

We start with the definition of the centroid curve for any cubic Hamiltonian H

having

a centre. It is easy to see that the

integral I(h,)

from

(0.2)

can be written in the form

where Int

(h)

is the

region

inside the

oval 6(h)

and = -

fx-gy.

Define the

integrals

Then the centroid

point [10]

of

region

Int

(h)

has coordinates

(~(h,), ~(ja))

where ~

= =

Y/M.

DEFINITION 1.1. - Let

(hl, h2~

be the maximal interval of existence of the oval

~i(h) (including

the

separatrix cycle

and the centre

inside).

Then

the curve

is called

[16]

the centroid curve of the

family

of ovals.

The

importance

of the

concept

of the centroid curve lies in the fact that its

geometry

contains the

complete

information about the

possible

zeros of

each

integral

=

aX(h)

+ +

(i.e.

for all values c~,

,~,

~y and all

perturbations f, g), although

the definition of L

depends only

on

H. We mention also the

following

obvious facts:

a)

L is affine

invariant; b)

the

endpoints

of L are the centroid Z of the area bounded

by

the

separatrix cycle

and the centre C

lying inside; c)

for

non-generic Hamiltonians,

L

is a line

segment; d) M ( h ) gives

the area of and T = M’ is

the

period

function. It has been

recently proved [8]

that T is monotone

and hence

M~~ ~

0.

Specifying

Definition 1.1 to the case of

symmetric

Hamiltonians

(0.3)

we

see that there are two continuous families of ovals

8 (h)

defined

respectively

in the intervals

~-h~,

and where 0

hs h~

are

given by

the formulas

Vol.

(7)

We shall denote

by Li

and

by L2

the

corresponding

centroid curves. We

point

out that in view of the

symmetry

of H the curve

L2

is

just Li

rotated

on an

angle

In the

sequel

we use L to denote any of

Ll, L2.

Further

denote

by Ci, C2

the centres and

by Sl , 82

the saddles of the Hamilto- nian

(0.3).

For later use we

compute

their

coordinates, obtaining

and

respectively

=

(xs, ~S) _ (-~c, -~~), C2

=

-G’~, ~‘2

=

The

following

theorem sums up several

properties

of the centroid curve

of

generic

Hamiltonians obtained in

[16]

and which we need in the

present

paper.

THEOREM 1.1. -

Suppose

that H is

generic.

Then

(i)

Near its

endpoints

the curve L is

regular,

i.e.

(h)

+

r~’2 (h) ~

0.

(ii)

Near the

endpoints of

L the curvature » is not zero and has the same

sign.

It is not hard to see

[ 16]

that for a

given perturbation f, g

in

(0.1)

with

( -I- ~,~ j ~ 0,

the number of limit

cycles

in

(0.1)

for small c is

equal

to the

number of the intersection

points (counting

the

multiplicities)

of L with the

line £ : ax +

-~- ~y

=

0, provided

L is

regular.

Moreover as c --~ 0 these

limit

cycles

tend to the ovals

8(h)

whose centroid

points

lie on £. Hence

our

goal

would be achieved if we show that

Li, L2

are

regular

and that any line intersects their union L =

L1

U

L2

in at most two

points.

In section 5

we prove the

regularity

of the centroid curves in the case of an

arbitrary generic

Hamiltonian with two centres,

using

Picard-Lefschetz

theory [3].

The

proof

is similar to that in

[16]

and

exploits

ideas from

[12]

and

[22].

Note that the

regularity

of L as well as the

nonvanishing

of the curvature

are affine invariant conditions. Most of our efforts will be concentrated to show that an

arbitrary

line £ intersects

Li

in no more than two

points.

Therefore the

following

theorem is the bulk of our construction:

THEOREM 1.2. - The curvature

of

the centroid curve

Ll

at each

point

is non-zero.

Our

plan

to show the

nonvanishing

of the curvature is to start with a

Hamiltonian H for which we

already

know that. Such one can be obtained for

example

from the standard

elliptic

Hamiltonian

Ho == y2 -~- x3 - ~

via small

perturbation

in the direction of the

symmetric

Hamiltonians. This is done in section 6. From Theorem 1.1 we know that in the

generic

case

the curvature near the

endpoints

of the centroid curve is

always

non-zero.

Annales de l ’lnstitut Henri Poincaré - Analyse non linéaire

(8)

This

implies

that

varying

the

parameters

in H if the curvature

acquires

a

zero this would be at least a double zero. It will

correspond

to a zero of

multiplicity

at least four of

I ( h) provided f, g

are

properly

chosen

([16], Corollary 2.1).

A

simple

but

important

fact is that the second derivative of I and the first derivative of the

period

function T

satisfy

Picard-Fuchs

system

of order two, and hence the ratio w =

I" /T’

satisfies a Riccati

equation.

This

observation,

used for the first time in

[ 11 ],

is based on the fact that the residua of the form w

= - ,~ dy + 9 dx

are linear in hand

therefore the second covariant derivative of the Gauss-Manin connection of w has no residua. In section 2 of the paper we use these ideas to derive Riccati

equation

for w. Most of the

present

paper is devoted to a

precise qualitative study

of the Riccati

equation

satisfied

by

w. In

particular

we

explore

the

properties

of the zero isocline of the

equation.

On its hand to

study

the zero isocline we include it in a

family

of curves

forming

the level

curves of a suitable Hamiltonian. We

repeat

the same

procedure

this time

considering

the horizontal and also the vertical isocline of the Hamiltonian

system

to obtain its

phase portrait.

From the

properties

of the zero isocline

of the Riccati

equation

we find the

possible

behaviour of w. The

upshot

is that w =

I" /T’ (and

hence

I ~~ )

has no double zeros. All this

analysis

is contained in section 4.

We have to

emphasize

that in section 4 we do not

study w

for all values

of the

parameters

a,

j3,

~y. We are interested

only

in values a,

corresponding

to lines f

tangent

to

Li.

In this case w is

subject

to

quite

restrictive

inequalities,

which we obtain in section 3. The

proof

of Theorem 1 is

completed

in section 7

by examining

the mutual

position

of the two centroid curves.

In

theory

the same

plan

could be

applied

to all another cases

(including

the one studied in

[16]). Unfortunately

the

computations

in

general

become

much more

complicated.

In

spite

of that we believe that the

plan

can be

performed, eventually

with some modifications. In this

respect

the

present

paper,

apart

of its own

importance,

can serve as a model for the

general

case.

2. PICARD-FUCHS SYSTEM AND RICCATI

EQUATION

In this section we derive several differential

equations

satisfied

by

Abelian

integrals.

Our final

goal

will be the Riccati

equation

satisfied

by I"

For this we need to derive Picard-Fuchs

system

for the

integrals Io, I_ 1, 1-2 ,

defined below.

(9)

Given h E

(-he, -hs),

we consider the level curve H =

xy2

+

3 x3 - x

=

h,

whose

equation

can be written also in the form

Put z = xy +

M/2

and define the

integrals

We have

and

similarly Y(h)

=

M(h) _ - I-1.

LEMMA 2.1. - The

integrals Io, I-1, ~-2 satisfy

the

following

system

of equations

Proof. - Express

the left and the

right

hand-side of the

identity

in terms of the functions

Ik, k

=

0, ~ 1, ...

This

gives

a connection

between these functions:

Similarly using

Annales de l’lnstitut Henri Poincaré - Analyse non linéaire

(10)

we obtain

Using (2.3)

with k = -1 and

(2.4)

with k = 0 we

get

the third

equation

of

(2.1).

In the same manner

applying (2.2)

and

(2.3)

with k = 0 and

(2.4)

with k == -1 we obtain the second

equation

of

(2.1).

At the

end,

if

we eliminate

Ii

and

1-3

from

(2.2), (2.3), (2.4)

with k

= -2,

we obtain the first

equation.

COROLLARY 2.2. - The

integrals X, Y,

M

satisfy

the system

Now we are in a

position

to derive the Riccati

equation

satisfied

by w(h)

= Put v = w - ~. Then we can consider that

v =

I~~/M~~

where in

I(h)

we

have ~

= 0. Differentiate once the first

two

equations

in

(2.5)

and twice the third one to

get

From these

equations

and from I = aX eliminate first Y and then X.

After

solving

with

respect

to

7

and

M

we obtain

Vol.

(11)

Here for shortness we have denoted

A i /

where

From

(2.6)

we derive the Riccati

equation

Here B =

B1

+

B2. Returning

to the variable w = v

-~ -y

and

writing (2.8)

as a

system

we

finally get

LEMMA 2.3. - The

following

Picard-Fuchs system is

satisfied:

this is our basic

system

which we will

study

in the next two sections. In

particular

of a

great importance

for our

analysis

will be the

properties

of the

zero isocline

Fo of (2.9), given by

the

points (h, w )

on the

algebraic

curve

(2.10)

We will often consider

Fo

as the

graph

of the two-valued function

wo (h)

determined

by P ( h, wo ( h) )

- 0

("+"

is

always assigned

to the upper

branch).

For later use denote

by Do

the discriminant of the

quadratic equation (2.10):

Annales de l’lnstitut Henri Poincaré - Analyse non linéaire

(12)

3. BASIC

INEQUALITIES

In this section we derive several

inequalities concerning

the

parameters

in

(2.9)

and the values of

wo

at the critical values of

H, provided

the line

P : ax +

{3y + "y

= 0 and the centroid curve

Li satisfy

some conditions

listed below. These

inequalities

are crucial in our

analysis

of

(2.9).

Through

this and the next section we suppose the

following:

(U) (i) ~

is a

tangent

to the centroid curve

Li

at an internal

point

~~~jL~, rl(h,~~~

(ii)

the curvature of

Li

does not

change

the

sign (but

may

vanish);

(13)

(iii) running L1,

£ rotates on an

angle

less than ~r.

We start with certain restrictions for the coefficients of the line £

imposed by (U).

LEMMA 3.1. - The

coefficients

a,

,~, ~y (multiplied if needed by -1) satisfy

the

inequalities

Proof. -

We are

going

to prove that the

angular

coefficient of £ satisfies the

inequalities

If this was done and

assuming

that

j3

> 0 we

immediately get

the last

two

inequalities

in

(3.1).

Moreover the

loop through Si

is

placed

in the

half-plane x

> 0 and it is

easily

verified that if

(~,

and

(.x, y2)

are

points

on

8 (h),

h E

(-h~. -ja5),

then ~/i + y2 0. Therefore

Li

is situated in the fourth

quadrant

and since 03B1 0

j3

the

equality

+

+ -y

= 0 is

possible only if 03B3

> 0.

In order to prove

(3.2)

we need another normal form

of the Hamiltonian. In these coordinates

Li corresponds

to values h E

[0, ~].

Having

in mind

(U),

the

proof

consists of

finding

the

equations

of the

tangents

at the

endpoints

of

Li. According

to

Corollary

3.4 from

[16]

the

equation

of the

tangent .~~

of

Li

at the centre

Cl(O,O)

-Ax

(see Fig. 2).

The

tangent .~s

at the other

endpoint

goes

through

the saddle

S 1

=

(1, 0)

and

through

the centroid of the

loop

area

Zi [16],

but the

coordinates of

Zi

are

given by extremely long

formulas. Because of this

we use another line instead of

.~s .

The

saddle-loop

is determined

by

the

equation

H =

~.

The line x

= 2

divides the

region

inside the saddle-

loop

into two

parts Ac

and

s containing respectively

the centre and the

saddle.

Obviously

the reflection of

As through

the line x

= 2

is contained in

Ac.

This means

that ~ ( 6 ) ~.

Then

using (U)

and results from

[16]

yields

that

Zi

=

(~ ( 6 ) , ~ ( 6 ) )

is located inside the

triangle

with vertices

( 0, 0 ) , ( 2 , 0 ) , ( 2 , - 2 ) .

Let m be the line

symmetric

to

.~~

with

respect

to the line x

= 2 ,

m has an

equation

y

= ~ ( x - 1).

Then the

angular

coefficient k of an

arbitrary tangent £

at an interior

point

of

Li

satisfies

Annales de l ’lnstitut Henri Poincaré - Analyse non linéaire

(14)

Fig. 2. _

the

inequalities -A 1~

A. In order to return to the coordinate

system

in which the normal form is

(0.3)

we

change

the variables

(x, y)

-

(xl,

where

In this coordinate

system

the

equations

of

.~~

and m become

respectively (we

omit the

subscript 1)

Hence the

angular

coefficient k of any

tangent

line satisfies

~ ~

1/~c.

Vol. 13, n°

(15)

As a direct consequence of

(2.7)

and

(3.1 )

we obtain

COROLLARY 3.2. - The constants

A; B, C, D,

E in

(2.9)

are

positive.

LEMMA 3.3. -

(i)

The values

ofw/ ( h)

at the ends

of the

interval

have the

following signs:

(ii)

The derivatives

satisfy

Proof. - (i) Using

the

explicit

formulas for

A, A,

B etc. we express the discriminant

Do

from

(2.11)

in the form

Also we have

This

gives

after obvious calculations

Annales de l ’lnstitut Henri Poincaré - Analyse non linéaire

(16)

Further,

with the

help

of

(2.7)

and the formulas for

~~

from section 1

we obtain via

straightforward computations

that

wo ( - ~~ ~

=

~x~ -f-,~3~~ -~-~y.

In the same way we

compute

At the end

where

From

(3.4)

it is clear that the

point ~~)

is

lying

on the

tangent .~~.

We will show further that

~~

xs.

Really

xs = -~~ and

~~

+

5~ 1 - I~2.

But x~ + ~/c

5h~

is

equivalent

to

The last

inequality easily

follows from

h~ 9 (2

+

3~) ~~ ( ~

+

~c) .

The

inequality ~~

shows that

~e)

is. above any

tangent

.~ to

Li

which

gives

that = +

+ ~

> 0. The other two

inequalities

in

(i)

follow from the fact that both the centre

~~)

and the

saddle

(xs,

are below any

tangent

.~.

(ii) Evaluating

the derivatives of

wo

at A = 0 we

get

Write

Do

in the form

(17)

This

gives

Calculation the values of A’

at - hs and -he yields:

Finally

we substitute h

= - h~

in the above formulas and via direct

computations

find

Repeating

the above

computation

for h

= - hs

we obtain

thus

proving

the lemma.

4. THE ZERO ISOCLINE OF THE RICCATI

EQUATION

In this section we list a number of

properties

of the zero isocline

Fo

of

(2.9) supposing (U)

hold. In

particular

decisive for our purposes will be the number and the mutual

positions

of the minima and maxima of

wo ( h) .

We start with the

following

lemma.

LEMMA 4.1. -

(i)

The curve

Fo is centrosymmetric

with respect to the

point (0,~).

..

(ii) Fo

has no compact components.

Proof -

Assertion

(i)

is obvious. To prove

(ii)

we observe that the discriminant

Do given by (2.11)

can have no more than two zeros for h 0.

Together

with the behaviour near h = 0

(see

the next

lemma),

this

eliminates the existence of

compact components.

Annales de l’lnstitut Henri Poincaré - Analyse non linéaire

(18)

Next we

study

certain

asymptotic properties

of h 0.

LEMMA 4.2. -

(i)

The

function wo (h)

has the

following asymptotic expansions

(ii)

The

function wo (h)

has the

following asymptotic expansions :

Proof. -

Direct

computations.

Our further

plan

is to consider the entire

family

of level curves

which is tantamount to the

phase portrait

of the Hamiltonian

system

This is because the discriminant

Do

is either

positive

or can

change

the

sign depending

on the

parameter

values in

(2.11 ). Consequently Fo

can bifurcate

and all

possible

situations should be occurred in the

family (Fp :

p E

I~~.

Again

for

simplicity

of the notation we can consider the

portrait for 7

=

0,

the

general

case

being

a mere translation w -~ w + ~y.

LEMMA 4.3. -

System (4.1)

has

exactly

one

singular point So

in the

half-plane h

0. The

point So

is a saddle.

Proof -

From the

equations Pw

=

Ph

= 0 eliminate w and

put h2

= z

to obtain the

equation

Obviously

this

equation

has at least one

positive

root zo. The

analysis

of

the coefficient alternations and Descartes’ criterion rules out the existence

Vol.

(19)

of other

positive

roots. As the Poincare index of the field in

(4.1)

cannot

exceed

0, using

the

symmetry yields

that

So

is a saddle

point.

Denote

by .ho == -~o

the abscissa of the saddle

So .

Our further

analysis

will concentrate

mainly

on

study

of two of the isoclines of

(4.1), namely

the horizontal

© ~

and the vertical one V

= ~ Pw

=

0 ~ .

The

properties

we need are easy to obtain and are summed up in the

following

lemma:

LEMMA 4.4. -

(i)

The curve V is concave in the

half-plane

h 0.

(iii)

The

asymptotics

hold

Proof -

Direct

computation.

’The horizontal isocline ?-~ is more

complicated

for

studying.

Moreover it bifurcates as

parameters

vary. To describe its

properties

we denote

by

~~~~~h),

two roots of the

equation Ph (h, w)

=

0,

h 0.

LEMMA 4.5. -

(i)

Let

~~ -

3AE > 0. Then consists

of

two separate

curves

given by

the

graphs of wH.(h),

h 0. Their

asymptotics

are the

following

Each

of

the

functions w~. ( h)

has

exactly

one maximum and no minima.

Let

.~2 -

3AE 0. Then the two

graphs of w~

meet at a

finite point

thus

fornfng geometrically

one smooth curve H. For the two

branches have the same

asymptotics

as in

(4.2).

The

function

has

always

one maximum. The

function w~ (h~

has either one minimum and one

maximum or no extrema, the latter case

including

also the

possibility of

a

collision

of

the maximum and minimum.

Proof. -

The

asymptotics

are

computed straightforward

from the formulas

Annales de l’Institut Henri Poincaré - Analyse non linéaire

(20)

When B2 - 3AE ~

0 both exist for all h 0. When

B2 - 3AE

0 the discriminant D has a

simple

zero at a

negative

value h =

h,

i.e.

=

wH(h)

and both are defined for h E

(-oo, h].

At the end

computing

the derivative of we obtain:

Putting h2

= z this

yields

a cubic

equation

for all extrema:

In the case

B2 -

3AE > 0 the

asymptotics (4.2)

say that each of has at least one maximum. As

(4.3)

has at most two

positive solutions,

this proves

(i).

In the case

B2 -

3AE 0 the

equation (4.3)

has either one or three

positive

solutions.

Using

the

asymptotics (4.2)

and the behaviour near h show that has at least one maximum while either has no

extrema or has one maximum and one minimum. To obtain more

precise information,

we observe that the derivative of

w~

is

negative

for z

jB/2

and the derivative of

~7~

is

positive

for z >

B / 2.

If

F ( z )

denotes the

polynomial

in

(4.3),

we have

F ( B ) _ - 3AE 2 0, F’ ( B ) _ - 8B E

0.

Hence

(4.3)

has

exactly

one zero

greater

than

B/2.

This

yields

that

wjj

has

always exactly

one extremum

(maximum).

Lemma 4.5 is

proved.

We will denote the two branches of H

by H+

and ?-~- .

Our next

objective

is to locate the saddle

So

with

respect

to the extrema of Before that we need more refine

study

of

H,

which we do

again by considering

the whole

family

of curves

Ph

= const and their horizontal and vertical isoclines

HH = ~ Ph h

=

4 ~

and

Hv

=

0 ~ .

Denote also

by

?-~C~~..r

the two branches of

HH (see Fig. 3).

As above we consider these curves as

graphs

of the

corresponding

functions

wH(h)

and

wHv(h), h

0.

LEMMA 4.6. -

(i)

The curves and

Hv

are concave. Each

of H±H

has

a maximum at =

6

and has a maximum at

hHV = B .

(ii)

The curves

HH

and

Hv

have no common

point

and

Hv

is situated

between the two branches

of HH.

Proof -

Direct

computation.

LEMMA 4.7. -

(i)

The curves V and

Hv

intersect at the

unique point of

maximum

of

V .

(21)

Fig. 3.

(ii)

The curves V and have no common

point,

V and

H-H

intersect at

a

unique point S*

with an abscissa

h* satisfying

(iii)

The saddle

So

is on the lower branch

of

H. Its abscissa

ho satisfies

Proof - (i)

is obvious.

(ii)

To prove the existence and

uniqueness

of

S*

we argue as above in

establishing

the existence of the saddle

So.

For the

proof

of the

inequality hHH h*

we first take the

asymptotic

values of

the functions

wHH(h)

and

wir (h)

when ~ --~

-0, namely

Annales de l’Institut Henri Poincaré - Analyse non linéaire

(22)

This shows that

wHH(h)

>

wv(h)

near h = -0. On the other hand we

can see that which in details reads

or

(3B2

+

36C)2 4B2(4B2

+

6ABD). Using

the

expressions

for

A, B, C, D

we

easily verify

that 12C

B2,

B AD. Then

(3Bz

+

36C)2 3684 4B2(4B2

+

SABD) 4B2(4B2

+

6ABD).

This proves

(4.4).

(iii)

We follow

essentially

the same

argument, showing

that

This condition is

equivalent

to

(B2 -~-12C)2 8B2(8B2 +

9ABD -

6AE)

which follows from 12C

B2

and the

easily

verified

inequality

AE

B2.

The

inequality (4.6)

we

proved implies

that V and the lower branch H- intersect at a

point So

with an abscissa

ho satisfying (4.5). Really

if

B2 -

3AE > 0 then

(4.2)

and Lemma 4.4

(iii) give

that >

wv (h)

near h = -0. If

B2 -

3AE 0 then

(4.6) implies

that

hHv h,

where h

is the zero of D from Lemma 4.5

(ii).

Hence =

wHv(h)

>

wv(h)

because h >

hHv

>

hv

and

wHv(h)

>

wv(h)

for h >

hv,

the latter a

consequence of

(i),

Lemma 4.4

(ii)

and Lemma 4.6

(i).

Thus

hHv ho h

and

(iii)

is

proved.

Another

property

of H and V we need is the

following

LEMMA 4.8. -

Suppose

that the

function wH (h)

has a maximum at

hH.

Then

hH ho (The

saddle

So

is to the

right of the

maximum

of w~(h)).

Proof. -

The extrema of H- lie on

~CH

and

~C~

itself is a concave curve with a maximum at a

point

with an abscissa

hHH.

Since

this

yields

that the maximum of ~C-

(if

it

exists)

has an abscissa

hH hHH

and moreover

wH (h)

for

h

Suppose

that

ho hH.

Then = and

= > because >

wv(h)

for h >

ho.

Therefore the intersection

point S*

of V and

H h

has an abscissa

h*

with

ho h* hH hHH -

a contradiction with

(4.4).

This proves the lemma.

In order to make the facts

proved

in Lemmas 4.3-4.8 more

applicable,

below we list all needed

points

of H and V with their coordinates and characteristics:

(23)

- the saddle

point, So

= H n

V,

the

point

of maximum of

~+, Sv(hv , wv (hv)) -

the

point

of maximum

of V,

the

point

of maximum of

~-,

the

point

of minimum of

~-C-,

~’(h, wH (h) ) -

the

point

where

H+

meets ~-~C’ .

(The

last three

points

do not exist

necessarily.)

Now we are able to describe in details the

picture

of the minima and maxima of

wo ( h ) .

LEMMA 4.9. -

(i)

Let

Do

>

0 for

any h. Then the

function

h 0

has

exactly

one maximum and one minimum.

(ii)

Let

Do

0

for h

E

(h-, h+),

h-

h+

0. Then

a) for

h E

( - oc , h - ~

the

function wo ( h )

has

exactly

one maximum and the

function wo {h)

has

exactly

one maximum and one

minimum; b) for

h E

~h+, 0)

the

function wo (h)

has

exactly

one maximum and one minimum.

Proof -

The

separatrices

H and V divide the half

plane h

0 into four open domains if

B2 -

3AE 0 and into five open domains otherwise. Let

us denote

by ~l~

and

Vi

the

parts

of ?~C- and V

respectively

which are to

the left of the saddle

So

and

by H2

and

V2 -

the rest

part

of H and V.

Analytically (for example)

we have

= ~ ( h, wH ( h) ) :

h

ho}

and

similarly

for the other arcs. In the case where

B2 -

3AE >

0, ~-C2

itself consists of two

separate

curves, one of them is

H+

and the other is the

part

of to the

right

of the saddle. We will write sometimes and

~-C2

in

order to

signify them,

thus

H2

=

H+2

n

?nC2

in this case. Denote

by Oij

the

domain in h 0 which

boundary

is

Hi

U

Vj, i, j

=

1, 2. (More precisely

in the case

B2 -

3AE > 0

03A922

consists of two

separate

subdomains:

5222

with

boundary H-2

U

V2

and

03A9+22

with

boundary U {h

=

0};

in the case

B2 -

3AE 0 the

boundary

of

f222

is

~C2

U

V2 U {h

=

0~ .)

Therefore

depending

on the

sign

of

B2 -

the

position

of the saddle

point

and the

availability

of extrema on ~-C ~ there are four cases I - IV

(see Fig. 4, a,b,c,d):

I

B2 -

3AE

0,

~C- has no extrema,

II

B2 -

3AE

0, So

is between

S~

and

S,

III

B2 -

3AE

0, So

is between

SH

and

,~~,

IV

B2 -

3.4E > 0.

Annales de l’Institut Henri Poincaré - Analyse non linéaire

Références

Documents relatifs

Actually, the use of normal forms in that framework has been introduced by Shatah [19] to prove global existence of solutions of non-linear Klein-Gordon equations on R 3 (an

and advantages to develop a system identification theory for Port-Controlled-Hamiltonian systems, as presented in chapter 1; the structural identifiability analysis

Bi-Hamiltonian systems on the dual of the Lie algebra of vector fields of the circle and periodic shallow water equations... ccsd-00102390, version 1 - 29

Section 3 contains a notion of a local quadratic growth condition of the Hamiltonian at the presence of control constraints and a formulation of generalized strengthened

In this section, we prove that any planar polynomial Hamiltonian system admitting centers can be perturbed along its gradient, so that each center bifurcates into a nest of

Par contre, dans notre étude, la mortalité a baissée en cas d’accouchement par voie basse instrumentalisée : 18 % de mortalité pour l’accouchement spontané contre

In 1984, Cohen and Lenstra [CoLe84] made a number of conjectures about the structure of the class group and divisibility properties of class numbers of real and imaginary

We shall find first the exact upper bound for the number of the zeros of an Abelian integral I(h) € As (we keep the notations of the preceding sections). Using the method of Petrov,